You are on page 1of 10

Superlattices and Microstructures 91 (2016) 391e400

Contents lists available at ScienceDirect

Superlattices and Microstructures


journal homepage: www.elsevier.com/locate/superlattices

Effects of an in-plane magnetic field on the energy dispersion,


spin texturing and conductance of double quantum wires
B. Gisi a, Y. Karaaslan a, S. Sakiroglu b, *, E. Kasapoglu c, H. Sari c, I. Sokmen b
a _
Physics Department, Graduate School of Natural and Applied Sciences, Dokuz Eylül University, 35390 Izmir, Turkey
b _
Physics Department, Faculty of Science, Dokuz Eylül University, 35390 Izmir, Turkey
c
Physics Department, Faculty of Science, Cumhuriyet University, 58140 Sivas, Turkey

a r t i c l e i n f o a b s t r a c t

Article history: We investigate the electronic structure, spin and transport properties of double quantum
Received 16 October 2015 wires formed by a symmetric, double quartic-well potential subjected to an in-plane
Received in revised form 21 December 2015 magnetic field by taking into account Rashba and Dresselhaus spin-orbit couplings. The
Accepted 22 December 2015
energy dispersion relation of the system is analyzed for different strengths of spin-orbit
Available online 24 December 2015
interactions, magnitude and direction of magnetic field. Our numerical results reveal
that the existence of aforementioned parameters modifies strongly the energy band
Keywords:
structure, introduces a wave vector dependence to energies and also leads to crossings and
Rashba and Dresselhaus coupling
In-plane magnetic field
anticrossings between subbands. This complex structure of energy dispersion gives rise to
Double quantum wire the appearance of square-wave like oscillations in the conductance. The spin-orbit cou-
Anharmonic potential plings, magnetic field, potential profile and Fermi energy of an electron significantly affect
the depth and width of conductance steps. Moreover, we found that the competing effect
between spin-orbit couplings and magnetic field leaves its marks on the spin texturing.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction

Recent advances in fabrication of high-mobility low-dimensional semiconductor structures have prompted the large
number of works in this area. The double low-dimensional semiconductor structures, such as double quantum wells, double
quantum wires and double quantum dots, constituted by restricting the motion of charge carriers are attractive phenomenon
in from various branches of physics to chemistry [1]. The one-dimensional double well potential is a well-known and topical
problem in quantum mechanics [2]. This potential consisting of two valleys separated by the potential barrier is utilized to
study the properties of numerous systems [3]. The study done by Bender et al. [4] for obtaining the energy eigenvalues of
quantum anharmonic oscillator has been quite impressive that has been followed by a great number of works. The energy
eigenvalues of double well anharmonic oscillator are investigated using perturbative or nonperturbative methods by many
researchers. Banerjee et al. [5] obtained the eigenvalues of two-well oscillator by a nonperturbative method WKB. Hodgson
et al. [6] performed an analytic procedure to calculate accurate wavefunctions and energy eigenvalues of a double minimum
potential with almost degenerate levels. The effect of an external field onto the behavior of the lowest eigenvalues of the
quantum double well anharmonic oscillator was surveyed by der Straeten et al. [7]. Also, Witwit [3] calculated energy levels of

* Corresponding author.
E-mail address: serpil.sakiroglu@deu.edu.tr (S. Sakiroglu).

http://dx.doi.org/10.1016/j.spmi.2015.12.032
0749-6036/© 2015 Elsevier Ltd. All rights reserved.
392 B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400

nonsymmetric double well potentials in one-, two- and three-dimensional systems for large values of the depth of the
potential.
Furthermore, a great deal of theoretical [8,9] and experimental [10,11] work has been devoted to survey the electronic and
transport properties of double quantum wires (QWs), consisting of two parallel long QWs coupled through a potential barrier
allowing the tunneling of electrons between them. Shi et al. [12] investigated the magnetotransport properties of dual QWs
for different heights of barrier in the presence of magnetic fields. Numerical results show that if the potential barrier is height
enough, i.e. the coupling between the QWs is weak so they are approximately independent, the lower energy eigenvalues are
closely bunched as a couple. Lyo and his collaborates [13,14] searched the energy spectrum, magnetization and magneto-
transport properties for tunnel-coupled double QWs subjected to an external magnetic field. The electronic transport
properties of double QWs by considering impurity, perpendicular magnetic field and correlated disorder were investigated by
Korepov [15,16]. Gudmundsson et al. [17] analyzed the energy spectrum, electronic transport and probability density of
electron waves of a double QW system containing a coupling element in the middle barrier between the two parallel QWs
under the influence of magnetic field.
The study of spin-orbit (SO) interaction in quantum confined structures is important from the standpoint of their
fundamental role in spintronic applications [18]. These devices are based on manipulating and controlling the spin orien-
tation by utilizing SO interaction which arises from inversion asymmetry properties of low-dimensional systems. Structural
inversion asymmetry induces Rashba SO interaction which can be tuned by changing the gate voltages [19] whereas bulk
inversion asymmetry gives rise to Dresselhaus SO interaction which contribution can be altered by sample thickness or
electron density [20]. Considerable amount of survey has been paid on the electronic, spin and transport properties of single
QWs taking into account the effects of external magnetic fields and SO interactions [21e29].
Band structure behaviors, spin texture and transport properties of double QWs have attracted immense interest in respect
of promising remarkable new devices, faster, smaller, and more powerful for the applications in quantum computing, spin-
field effect transistors and spin filters [30,31]. It is therefore important to characterize electronic energy spectrum of double
QWs in the presence of different external agents, in order to gain a better understanding of the conductance and spin-related
properties. In this purpose Karaaslan et al. [32] investigated the effects of Rashba SO interaction on the electronic energy
dispersion and conductance of double QW under the influence of perpendicular magnetic field. Recently some experimental
techniques have been developed to control a coupling of spin to electric field [33,34]. According to the result of these ex-
periments, an in-plane magnetic field is more appropriate to obtain a prominent spin resonance. SO interactions generate an
effective in-plane magnetic field which gives rise to an drift-driven in-plane spin polarization, and the summation of an
external magnetic field and SO induced effective magnetic field leads to observe surprising spin polarization [33,35]. In the
light of these findings it is worth to investigate the effect of an in-plane magnetic field on the energy, spin texture and
conductance of double QW considering the SO interaction. The organization of the paper is as follows: In Section 2 we briefly
present the system and methodology used throughout our study. In Section 3 we discuss the numerical results while Section
4 summarizes our findings.

2. Theory and formalism

Semiconductor nanowires are powerful class of materials that through a controlled growth and organization can be
produced in a wide range with a compositional and/or doping modulation [36]. Spatially separated two identical quasi-one-
dimensional systems can be coupled by an additional lateral confinement to produce a double quantum wire structure [37].
Therefore, the confinement potential representing the local depletion of stacked two-dimensional gases contained in double-
quantum-well heterostructures with controllable tunneling barrier thickness can be modeled by a double quartic-well
confinement potential given as

 2
1 m2
VðyÞ ¼ l y2  ; (1)
4 l

where m and l are positive, adjustable structural parameters controlling the height of the barrier between wells and the width
of wells. In-plane magnetic field applied with an arbitrary orientation is chosen as B ¼ Bðcosfex þ sinfey Þ where f stands as
an azimuthal angle. The wire orientation along x direction is shown in Fig. 1.
The solutions of Schro€dinger's equation can be written as a product of plane waves propagating along the x axis and y
dependent spinor part as follows:
 
eikx x 4nkx ðy; [Þ
jnkx ðrÞ ¼ pffiffiffi : (2)
L 4nkx ðy; YÞ

Here 4nkx is the spinor function. We use integer numbers n ¼ 1,2,3,… to label different energy subbands and introduce a
continuous wave number kx. The finite-T frame formalism is used as a numerical trick, but results of actual calculations are for
T ¼ 0 case. sa(a ¼ x,y,z) being the corresponding Pauli matrix, we can calculate spin magnetization components,
B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400 393

Fig. 1. The model of the double quantum wire structure and orientation of magnetic field (color online).

X L Z
ma ðyÞ ¼ dk〈jnk jdðri  rÞsa jjnk 〉ri fb ðεnk ; mÞ; (3)
n 2p

where fb ðεnkx Þ is the Fermi distribution.


The single-particle Hamiltonian of the quasi-one-dimensional wire system is given by

H ¼ H 0 þ H Z þ H R þ H D: (4)

First term H 0 , consists of the kinetic and confinement potential whereas the second term H Z describes the Zeeman effect
contribution arising from an in-plane magnetic field that are given explicitly:

p2x þ p2y
H0 ¼ þ VðyÞ; (5)
2m
 
H Z ¼ g  mB B cosfsx þ sinfsy 2: (6)

In the above expressions, px and py are the actual components of linear momentum, m* is effective mass, g* and mB are the
effective Lande-g factor and Bohr magneton, respectively. sx and sy represent the Pauli spin matrix components.
The last two terms in Eq. (4) are spin-orbit contributions where Rashba Hamiltonian ðH R Þ is given as

aR 
HR¼ py sx  px sy ; (7)
Z
and Dresselhaus Hamiltonian ðH D Þ is

aD 
HD ¼ px sx  py sy : (8)
Z

aR and aD are Rashba and Dresselhaus SO coupling parameters, respectively. We consider double quantum wire system
defined in narrow-gap material (InAs) because of its large spin-orbit couplings [38,39]. Despite the Rashba SO coupling effect
is found to be very pronounced in these semiconductor layers, our particular interest is exploring the system exhibiting both
sizable Rashba and Dresselhaus SO couplings. The Dresselhaus SO interaction in semiconductor quantum wires shows a
strong dependence on the geometry of the confinement and is particularly dominant in structures with strong and symmetric
confinement potential [25]. The bulk inversion asymmetry-induced effects on the spin-split energy levels and spin polari-
zations in QWs can be tracked for linear, quadratic and cubic terms (has particularly significant effects on the excited states)
[40]. However, the most important contribution to spin-polarized states comes from the linear Dresselhaus SO interaction
which is always present in systems with a zinc-blende structure (irrespective of any other spin-orbit interaction). Therefore,
the symmetric confinement potential adopted in our study renders possible the assumption of the k-linear contribution of the
Dresselhaus SO interaction.
The energy eigenvalues and eigenfunctions of Schro €dinger equation are obtained by using the finite element method
(FEM) in one-dimension which is based on expressing the wave function as a linear combination of interpolation polynomials
multiplied by as-yet-unknown coefficients in each of these elements [41e43]. The scaled form of the Hamiltonian can be
written as:
394 B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400

1  
H ¼ s p2 þ hR sx  sy hD pey
2 0 ey
 
1 2 1e 4 1 2 2 1 m e4
þ K 0 þ ley  m e ey þ s0
2 4 2 4 el
   
1 1
þ B cosf þ hD K 0 sx þ B sinf  hR K 0 sy :
2 2

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Z=m u0 ) is taken as a length scale and correspondingly
Characteristic length corresponds to harmonic oscillator (l0 ¼
the obtained energies are in Zu0 units. And also, we define dimensionless parameters: K 0 ¼ kx l0 , e y ¼ y=l0 , pey ¼ 1i v .
ve
y
pffiffiffiffiffiffiffiffiffiffiffiffi
 e  2
Throughout our work we use dimensionless form of the structural parameters: m e ¼ m= m u0 and l ¼ lZu0 =ðm u0 Þ . Lastly
2

B is proportional to Lande-g factor and the strength of magnetic field (B ¼ ðeg  BÞ=ð2m0 u0 ), e and m0 is the charge and the
mass of electron, respectively).

3. Results

The system under investigation is assumed to be InAs semiconductor heterostructure with corresponding bulk parame-
ters: g* ¼ 15, m* ¼ 0.03m0. The experimental values of SO coupling constants aR,D are in order of 1011 eV m [44,45]. We set
Zu0 to 2 meV and for the sake of simplicity, one of the structural parameters controlling the shape of the double-well potential
is fixed, e
l ¼ 1. In the present work characterization of the SO regime is realized by the ratio of the strength of SO interaction to
the confinement potential energy, DRðDÞ ¼ m a2RðDÞ =2Z3 u0 . By virtue of this fact, parameters DRðDÞ a0:1 define the strong SO
regime and DR(D) < 0.1 accounts for the weak SO regime.
In this study, we have theoretically investigated the energy spectra of double QW for several cases including different
strengths of SO coupling and magnitude and direction of magnetic field. We have aimed to reveal the qualitative hallmarks of
SO interaction and in-plane magnetic field in double QW so we present some illustrative examples rather than study whole
parameter space. Initially, in order to identify how the interplay between Rashba and Dresselhaus SO coupling affects the
energy subband structure of double QW, in Fig. 2 we give Ek graphs for different strengths of SO contributions in the absence
of magnetic field. We consider strong coupling limit, namely m e ¼ 1, in which tunneling effect between QWs is in action. In all
subfigures, the dashed lines represent the energy spectra of double QW with no SO interaction and magnetic field. It is clear
from the Fig. 2 that all subbands are two-fold degenerate for spin-up and spin-down states for all values of kxl0 presented with
dashed lines. Taking into account SO interaction lifts the spin degeneracy of the subbands for nonzero wave vectors and
produces an energy splitting proportional to kx. For strong Rashba SO coupling (DR ¼ 0.2) and weak Dresselhaus SO coupling

Fig. 2. Subband energy spectra as a function of linear momentum kxl0 for different values of Rashba/Dresselhaus SO couplings in the absence of magnetic field.
We set me ¼ 1. (a) DR ¼ 0.2 and DD ¼ 0.01, (b) DR ¼ 0.2 and DD ¼ 0.1, (c) DR ¼ DD ¼ 0.01, (d) DR ¼ DD ¼ 0.1. The dashed lines correspond the case of DR ¼ DD ¼ 0 and
B ¼ 0 (color online).
B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400 395

(DD ¼ 0.01) nearly parabolic structure as seen in Fig. 2(a). More complex subband structure emerges in the case of simul-
taneous contribution of both of SO interaction terms in strong regime (DR ¼ 0.2 and DD ¼ 0.1) (Fig. 2(b)). We observe local
extrema that can be attributed as an anomalous plateaus near kxl0 which plays an important role in the explanation of
anomalous steps in conductance. In the absence of magnetic field for equal strengths of Rashba and Dresselhaus SO in-
teractions the energy dispersion exhibits an interesting feature. In Fig. 2(c), the weak SO coupling (DR ¼ DD ¼ 0.01) generates
two horizontally displaced parabolas, which are uniformly shifted downward compared to the spin-degenerate band (the
dashed lines). Enhancement in SO interaction results in stronger deviation from parabolicity of the subbands and downward
shifting of energy as seen in Fig. 2(d). Also crossings between same levels with the opposite spins and anticrossings between
different levels appear.
We examine the influence of an in-plane magnetic field on the subband structure, because it has an important role to
clarify the beating pattern in the magnetoresistance and absorption spectra of magnetoeoptic transitions. The magnetic field
lifts the degeneracy of the two spin states via Zeeman effect and changes the character of the energy spectrum as can be seen
in Fig. 3. We have obtained that, when SO interaction is considered the zero k spin-splitting energy is not equal to the
Zeeman spin-splitting given by g*mBB, it also depends on the strength of SO interaction. Coexistence of SO interaction and
magnetic field gives rise to peculiarities in the energy spectrum and constitutes the basis of many spintronic devices
[18,46,47]. We searched the energy spectra of double QW for four different orientations (f ¼ 0; p=6; p=3; p=2) of a constant
magnetic field of 3 T for the strong SO regime characterized by DR ¼ 0.05 and DD ¼ 0.2. When magnetic field is not considered,
the SO interaction gives rise to symmetric double minimum structures in the energy subband which is shown by dashed lines,
but including the magnetic field breaks down this symmetry and introduces additional complexities in the subband structure
as shown in Fig. 3. The induced asymmetry can be explain via Eq. (9). For a magnetic field directed along the wire axis, cosf ¼
1 and B cosf contributes negative values to Hamiltonian for all kx. Also the Dresselhaus SO term is negative for kx < 0, and
positive for kx > 0. For negative (positive) kx values, due to the same (opposite) signs of Zeeman and Dresselhaus SO inter-
action, their overall effect contributes to the Hamiltonian. This explains why the minimum point of energy subband occurs at
negative kx as seen in Fig. 3(a). In the case of magnetic field orientated along the perpendicular axis of wire (f ¼ p=2), B sinf
has a nontrivial negative value. For positive (negative) kx values, due to the same (opposite) signs of Rashba SO interaction and
magnetic field, their overall effect contributes to the Hamiltonian. This elucidate the shifting of the minimum of energy to
positive kx as observed in Fig. 3(d). For the other orientations of magnetic field such as f ¼ p=3; f=6 (Fig. 3 (b) and (c)) the
asymmetries in the subbands can be explained in a similar manner. Additionally, the crossings and anticrossings arising from
SO interaction are observed as it is evident from the Fig. 3.
The energy spectrum of electrons critically depends on the effective geometry of QW as well as the external and internal
parameters such as magnetic field and SO interactions. In order to show the effect of structural parameters, in Fig. 4 we
present the energy dispersion behavior of the subbands of the double QW comprising of strong SO interaction (DR ¼ 0.2 and
DD ¼ 0.05) and magnetic field (1 T) orientated with an angle f ¼ p/4. The dashed lines correspond to the pure case (no SO

Fig. 3. Variation of the energy dispersion curves of electrons in a double QW for different orientations of magnetic field: (a) f ¼ 0, (b) f ¼ p/6, (c) f ¼ p/3, and (d)
f ¼ p/2. The magnitude of magnetic field and the strength of SO interactions are fixed at B ¼ 3 T, DR ¼ 0.05 and DD ¼ 0.2. The dashed lines correspond the case of
B ¼ 0 (color online).
396 B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400

Fig. 4. Energy spectrum of double QW in the presence of strong SO interaction (DR ¼ 0.2, DD ¼ 0.05) and an external magnetic field (B ¼ 1 T). The orientation
e are taken: (a) 1, (b) 1.5, (c) 2, and (d) 2.5. The dashed lines correspond the case of DR ¼ DD ¼ 0 and B ¼ 0
angle is chosen as f ¼ p=4 where different values of m
(color online).

interaction and no magnetic field) in all subfigures. The most striking feature of the figure is that the energy band levels shift
to upper values and are grouped in pairs by starting from the lowest level with increasing values of m e for both solid lines and
dashed lines as depicted in the figure. Small values of m e correspond the strong coupling limit (Fig. 4(a), (b)) in which the
tunneling effect between wires is allowed. In this limit, energy levels have local minimum and maximum points which give
rise to the appearance of intriguing properties in conductance. On the other hand, in the weak coupling limit (Fig. 4(c), (d)),
characterized by large values of m e, tunneling of electrons can be negligible and also the wires can be considered as two in-
dividual QWs. The energy spectrum of this decoupled double QW is a superposition of the dispersion curves of individual
QWs. The crossings and anticrossings between energy levels which appear in the strong coupling limit vanish in the weak
coupling limit. Further the energy subbands exhibit a smoother behavior.
When SO coupling is taken into account, spin is not a good quantum number and this results in occurrence of spin textures
across the wire with the spin direction depending on kx and the wire transversal coordinate y. Contrarily, when SO interaction
is not included, the states are exact eigenspinors even in the presence of in-plane magnetic field. Thus, observation of a clear
spin texture can be interpreted as a signature of SO interactions. Whenever the typical SO interaction energy scale becomes
comparable to the subband splitting, emergence of a sizable spin z component due to anticrossings between neighboring
subbands precludes a definition of common spin-quantization axis [22]. The combined effect of external magnetic field and
Rashba and Dresselhaus SO interactions has important role on the spin-related properties. Zeeman effect tends to align the
spin along the external magnetic field, whereas the SO interaction is responsible for the randomization of the spin direction.
The interplay between SO coupling and Zeeman effect leads to observation a variety of physical phenomena. The concept of
spin texturing is crucial for spintronics in two respects: it is closely related to that of spin current which is measured as a
change in the local spin and it represents the spatial distribution of the effective magnetic field due to spin-orbit interaction.
In the following part of the study, we illustrate to what extend effects of SO interaction and magnetic field modify the spin
orientation. In this respect we calculated spin expectation values 〈sn 〉 ¼ Jynk sn Jnk where n ¼ x,y,z. In-plane spin distributions
for the lowest subband at three different momenta are shown in the upper panel of Fig. 5, in the presence of strong SO
coupling characterized by DR ¼ 0.05 and DD ¼ 0.2 and an external magnetic field (B ¼ 3 T) orientated along the wire axis f ¼ 0.
The vector plot shows the in-plane spin whereas the solid line corresponds to the z component. The electrons in the double
QW experience three different magnetic fields: an in-plane external magnetic field, effective pseudomagnetic fields arising
!
from Rashba and Dresselhaus SO couplings. Effective Rashba magnetic field ð B R Þ is directed along the y axis because it is
!
perpendicular to the electric field and the electron's velocity [24]. The effective Dresselhaus magnetic field ð B D Þ acts along the
! !
x direction [25]. Electrons with positive (negative) kx experience þy (y) directed B R and x (þx) directed B D . Therefore the
net magnetic field exerted on electrons is the vector sum of the pseudomagnetic fields and the external magnetic field. In the
absence of SO coupling, the spins are parallel to the direction of external magnetic field. In the presence of SO interaction
effects, when f ¼ 0, for kx ¼ 0.69 (the minimum point of the lowest energy level) electron is under the influence of y
! !
directed B R , þx directed B D and þx directed external magnetic field. Similarly, when kx ¼ 0.69 (the symmetry point
B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400 397

Fig. 5. Upper panel displays the spin texture for three different propagation momenta for the values of Fig. 3(a). Plots from left to right correspond to momentum
kxl0 ¼ 0.69,0.0,0.69, where arrow length is proportional to spin density. Red solid lines correspond to the z component of magnetization. Lower panel presents
spin projections of eigenstates in the lowest spin-split subbands. The solid (dashed) lines show first (second) level of the lowest spin-split subband. Blue lines
indicate 〈sx〉 whereas red lines represent 〈sy〉 (color online).

!
according to kx ¼ 0 of the minimum point of the lowest energy level) electron is under the influence of þy directed B R , x
!
directed B D and þx directed external magnetic field. And also we should note that the effect of pseudomagnetic fields at
kx ¼ 0 is seen in the mid panel. Therefore as seen in the upper panel of Fig. 5 the spin orientation is directed along the effective
net magnetic field. Lower panel shows the dependence on kx of the spin expectation values in the first (solid line) and second
(dashed line) level of the lowest spin-split subbands where blue (red) lines indicate 〈sx〉 (〈sy〉). It can be seen that the spins of
eigenstates with large absolute value of kx approach to asymptotic states of 〈sx〉 and 〈sy〉. For negative large kx the spinor is
nearly described by eigenstates of 〈sx〉. Also, it is observed from the figure that, there exists right (left) moving states for both
spin-up and spin-down electrons at any energy.
In Fig. 6, we present the spin texture and spin expectation values for the same parameters with Fig. 5 but the orientation
angle of magnetic field is p/4. The spin orientations can be interpreted similarly. Due to the combined effect of pseudo-
magnetic field and y component of the external in-plane magnetic field, the spins turn to positive y direction which differs
from Fig. 5. Especially for electrons with positive kx experience both effective Rashba and external magnetic field in the þy
direction and the combination of these two fields leads to align spins nearly parallel to y axis. Similar behavior for the spin
expectation values as for f ¼ 0 is seen in the lower panel of Fig. 6, but with a smoother variation for the second level of the
lowest spin-split subband.
In low-dimensional electron systems, the transmission of electrons can be attributed as the redistribution of incoming
electron flux between the discrete energy eigenvalues of the system. Hence the determination of the energy dispersion of the
system is extremely important for the calculation of ballistic quantum conductance. Thus the total conductance of the system
is completely determined from the energy spectrum and Fermi energy. We assume that two reservoirs are connected to
double QW and the wire is long enough so that is devoid of backscattering process. Conductivity of the system, by neglecting
the electroneelectron interaction, can be simply calculated using Landauer-Büttiker formalism [48,49].
X
G ¼ G0 Taa0 (9)
aa0

where Taa0 is the transition probability from ja〉 state to ja0 〉 state and G0 ¼ 2e2/h is the conductance quantization. Following
the strategy of Ref. [50], the conductance can be calculated as:

e2 X X ðn;sÞ  n;s
G¼ b f Ei : (10)
h n;s i i
398 B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400

Fig. 6. Same with Fig. 5 for f ¼ p=4 (color online).

Here n and s represents the level of state and the level of spin, respectively while Ein;s is the energy of ith extremum point in
ðn;sÞ
the related energy subband and f ðEin;s Þ is the Fermi-Dirac distribution. bi corresponds 1 for a maximum and þ1 for a
minimum point in the energy subband labeled with n and s.
In this part of the paper, our goal is to calculate zero-temperature conductance for different values of structural parameter
involving the interplay between SO interaction and in-plane magnetic field. In the pure case, namely no magnetic field and no
SO coupling, the energy subbands are two-fold spin-degenerate for the strong coupling limit as seen in Fig. 4(a), and (b)
plotted with dashed lines. When the Fermi energy of an electron increases, the conductance increases 2(e2/h) instead of 1(e2/
h) as given in Fig. 7(a), (b) with red lines. The underlying physical reason can be explained as the additional propagating

e: (a) 1, (b) 1.5, (c) 2, (d) 2.5. Red lines correspond the
Fig. 7. The conductance as a function of Fermi energy for the parameters in Fig. 4 for varying values of m
conductance of double QW in the case of DR ¼ DD ¼ 0 and B ¼ 0 (color online).
B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400 399

channels which contribute to conductance are opened in pairs. For m e ¼ 2 (Fig. 7 (c)) the spin-degenerate lowest levels overlap
which leads to an increment of conductance in amount of 4(e2/h). In the weaker coupling limit, all of spin-degenerate levels
are grouped in pairs so conductivity increases monotonically with 4(e2/h) as is evident from the Fig. 7(d) with red line. For all
subfigures it is clear that with the increment of m e the energy gaps raise and this leads to observe wider plateaus of
conductance. The conductance under the influence of both magnetic field (B ¼ 1 T) with the orientation of f ¼ p=4 and strong
SO coupling limit (DR ¼ 0.2, DR ¼ 0.05) is given with blue lines. The competing effect of SO coupling and external magnetic
field gives rise to local minimum (subband bottom) and local maximum (subband top) in the subband structures. The ex-
istence of these minimum and maximum is the reason of anomalous steps in the conductance that appear on the top of the
ordinary steps. In this case, the conductance is anymore a stepwise monotonically increasing function, has dips at the
extremum points of energy level. The profile of conductance behaves like square-wave oscillations for which, with the
increment of me the width of dips become wider.

4. Conclusion

We discussed the energy spectrum, spin texturing, and zero-temperature ballistic conductance of double quantum wires
formed by nonparabolically confinement of a two-dimensional electron gas taking into account the influence of Rashba and
Dresselhaus spin-orbit interactions and in-plane magnetic field. We show that the combined effect of SO coupling and
magnetic field leads to complicated energy spectra and affects significantly the spin texturing which constitutes the basis of
spintronic devices. Moreover, we investigate the effect of confining potential on to the energy spectra and transport prop-
erties. We obtained that the profile of conductance behave like square-wave oscillations. The origin of the conductance
oscillation is the local extremas in the subband structure. The depth and the width of the square conductance strongly depend
on the strength of Rashba and Dresselhaus spin-orbit interactions, magnitude and direction of magnetic field, height and
width of the double well potential. Briefly, we found that the interplay of different SO couplings and magnetic field brings out
the modification of the energy dispersion and consequently the conductance and phase shift between the modulations of the
spin density components. Electrostatically modulated spin-orbit interaction and external magnetic field in a quantum wire
can be gainfully employed for the future design of spin-field effect transistors and spin filters, which rely on coherent spin
transport.

Acknowledgment

The authors B. Gisi and Y. Karaaslan would like to thank the Scientific and Technological Research Council of Turkey
_
(TÜBITAK _
BIDEB 2211) for doctoral scholarship. This work has been completed at the Dokuz Eylül University, Graduate School
of Natural and Applied Sciences and is the subject of the forthcoming Ph.D. Thesis of Bircan Gisi.

References

[1] P. Pedram, M. Mirzaei, S.S. Gousheh, Mol. Phys. 108 (2010) 1949.
[2] S.K. Bhattacharya, Phys. Rev. A 31 (1985) 1991.
[3] M.R.M. Witwit, J. Comput. Phys. 123 (1996) 369.
[4] C.M. Bender, T.T. Wu, Phys. Rev. D 7 (1973) 1620.
[5] K. Banerjee, S.P. Bhatnagar, Phys. Rev. D. 18 (1978) 4767.
[6] R.J.W. Hodgson, Y.P. Varshni, J. Phys. A Math. Gen. 22 (1989) 61.
[7] E.V. der Straeten, J. Naudts, J. Phys. A Math. Gen. 39 (2006) 933.
[8] K.J. Thomas, J.T. Nicholls, M.Y. Simmons, W.R. Tribe, A.G. Davies, M. Pepper, Phys. Rev. B 59 (1999) 12252.
[9] D.W. Wang, E.G. Mishchenko, E. Demler, Phys. Rev. Lett. 95 (2005) 086802.
[10] J.S. Moon, M.A. Blount, J.A. Simmons, J.R. Wendt, S.K. Lyo, J.L. Reno, Phys. Rev. B 60 (1999) 11530.
[11] M. Yamamoto, Y. Tokura, Y. Hirayama, M. Stopa, K. Ono, S. Tarucha, AIP Conf. Proc. 772 (2005) 925.
[12] J.R. Shi, B.Y. Gu, Phys. Rev. B 55 (1997) 9941.
[13] S.K. Lyo, D. Huang, Phys. Rev. B 64 (2001) 115320.
[14] D. Huang, S.K. Lyo, K.J. Thomas, M. Pepper, Phys. Rev. B 77 (2008) 085320.
[15] S.V. Korepov, M.A. Liberman, Phys. Rev. B 60 (1999) 13770.
[16] S.V. Korepov, M.A. Liberman, Phys. B 322 (2002) 92.
[17] V. Gudmundsson, C.S. Tang, Phys. Rev. B 74 (2006) 125302.
[18] S. Datta, B. Das, Appl. Phys. Lett. 56 (1990) 665.
[19] E.I. Rashba, Sov. Phys. Solid. State 2 (1960) 1109.
[20] G. Dresselhaus, Phys. Rev. 100 (1955) 580.
[21] J. Knobbe, Th. Scha €pers, Phys. Rev. B 71 (2005) 035311.
[22] M. Governale, U. Zülicke, Solid State Commun. 131 (2004) 581.
[23] S. Zhang, R. Liang, E. Zhang, L. Zhang, Y. Liu, Phys. Rev. B 73 (2006) 155316.
[24] P. Upadhyaya, S. Pramanik, S. Bandyopadhyay, Phys. Rev. B 77 (2008) 045306.
[25] S. Gujarathi, K.M. Alam, S. Pramanik, Phys. Rev. B 85 (2012) 045413.
[26] L. Serra, D. Sanchez, R. Lopez, Phys. Rev. B 72 (2005) 235309.
[27] F. Malet, M. Pi, M. Barranco, L. Serra, E. Lipparini, Phys. Rev. B 76 (2007) 115306.
[28] C.S. Tang, S.Y. Chang, J.C. Cheng, Phys. Rev. B 86 (2012) 125321.
[29] S. Sarıkurt, S. Ṣakiroglu, K. Akgüngo _ So
€ r, I. €kmen, Chin. Phys. B 23 (2014) 017012.
[30] P. Upadhyaya, S. Pramanik, S. Bandyopadhyay, Phys. Rev. B 77 (2008) 155439.
[31] M. Scheid, M. Kohda, Y. Kunihashi, K. Richter, J. Nitta, Phys. Rev. Lett. 101 (2008) 266401.
[32] Y. Karaaslan, B. Gisi, S. Sakiroglu, E. Kasapoglu, H. Sari, I. Sokmen, Supperlattice Microstruct. 85 (2015) 401.
400 B. Gisi et al. / Superlattices and Microstructures 91 (2016) 391e400

[33] Y. Kato, R.C. Myers, D.C. Driscoll, A.C. Gossard, J. Levy, D.D. Awschalom, Science 299 (2003) 1201. Y. Kato, R. C. Myers, A. C. Gossard, D. D. Awschalom,
Phys. Rev. Lett. 93 (2004) 176601.
[34] E.A. Larid, C. Barthel, E.I. Rashba, C.M. Marcus, M.P. Hanson, A.C. Gossard, Phys. Rev. Lett. 99 (2007) 246601.
[35] V.M. Edelstein, Solid State Commun. 73 (1990) 233.
[36] Y. Li, F. Qian, J. Xiang, C.M. Lieber, Mater. Today 9 (2006) 18e27.
[37] S.F. Fischer, G. Apetrii, U. Kunze, D. Schuh, G. Abstreiter, Phys. Rev. B 74 (2006) 115324.
[38] T. Meng, J. Klinovaja, D. Loss, Phys. Rev. B 89 (2014) 205133.
[39] E. Papp, C. Micu, D. Racolta, AIP Conf. Proc. 1564 (2013) 90.
[40] L. Villegas-Lelovsky, C. Trallero-Giner, M. Rebello Sousa Dias, V. Lopez-Richard, G.E. Marques, Phys. Rev. B 79 (2009) 155306.
[41] O.C. Zienkiewicz, R.L. Taylor, J.Z. Zhu, Finite Element Method: its Basis and Fundamentals, sixth ed., Elsevier Butterworth-Heinemann, 2005.
[42] I. Babuska, J.E. Osborn, Math. Comput. 52 (1989) 275.
[43] J.E. Pask, B.M. Klein, P.A. Sterne, C.Y. Fong, Comp. Phys. Commun. 135 (2001) 1.
[44] S. Pramanik, S. Bandyopadhyay, M. Cahay, Phys. Rev. B 76 (2007) 155325.
[45] J. Ko€nemann, R.J. Haug, D.K. Maude, V.I. Fal’ko, Phys. Rev. Lett. 94 (2005) 226404.
[46] S. Bandyopadhyay, M. Cahay, Appl. Phys. Lett. 85 (2004) 1814.
[47] S.D. Ganichev, E.L. Ivchenko, V.V. Bel’kov, S.A. Tarasenko, M. Sollinger, D. Weiss, W. Wegscheider, W. Prettl, Nature (Lond.) 417 (2002) 153.
[48] R. Landauer, IBM J. Res. Dev. 1 (1957) 223. Phys. Rev. A 85 (1981) 91.
[49] M. Büttiker, Y. Imry, R. Landauer, S. Pinhas, Phy. Rev. B 31 (1985) 6207.
[50] Y.V. Pershin, J.A. Nesteroff, V. Privman, Phys. Rev. B 69 (2004), 121306(R).

You might also like